Artificial Neural Networks

advertisement
Artificial Neural Networks
Chung-Ming Kuan
Institute of Economics
Academia Sinica
Abstract
Artificial neural networks (ANNs) constitute a class of flexible nonlinear models designed to mimic biological neural systems. In this entry, we introduce ANN using familiar
econometric terminology and provide an overview of ANN modeling approach and its implementation methods.
† Correspondence: Chung-Ming Kuan, Institute of Economics, Academia Sinica, 128 Academia Road,
Sec. 2, Taipei 115, Taiwan; ckuan@econ.sinica.edu.tw.
†† I would like to express my sincere gratitude to the editor, Professor Steven Durlauf, for his patience
and constructive comments on early drafts of this entry. I also thank Shih-Hsun Hsu and Yu-Lieh Huang
for very helpful suggestions. The remaining errors are all mine.
1
Introduction
Artificial neural networks (ANNs) constitute a class of flexible nonlinear models designed
to mimic biological neural systems. Typically, a biological neural system consists of
several layers, each with a large number of neural units (neurons) that can process the
information in a parallel manner. The models with these features are known as ANN
models. Such models can be traced back to the simple input-output model of McCulloch
and Pitts (1943) and the “perceptron” of Rosenblatt (1958). The early yet simple ANN
models, however, did not receive much attention because of their limited applicability
and also because of the limitation of computing capacity at that time. In seminal works,
Rumelhart et al. (1986) and McClelland et al. (1986) presented the new developments of
ANN, including more complex and flexible ANN structures and a new network learning
method. Since then, ANN has become a rapidly growing research area.
As far as model specification is concerned, ANN has a multi-layer structure such that
the middle layer is built upon many simple nonlinear functions that play the role of neurons in a biological system. By allowing the number of these simple functions to increase
indefinitely, a multi-layered ANN is capable of approximating a large class of functions
to any desired degree of accuracy, as shown in, e.g., Cybenko (1989), Funahashi (1989),
Hornik, Stinchcombe and White (1989, 1990), and Hornik (1991, 1993). From an econometric perspective, ANN can be applied to approximate the unknown conditional mean
(median, quantile) function of the variable of interest without suffering from the problem
of model misspecification, unlike parametric models commonly used in empirical studies. Although nonparametric methods, such as series and polynomial approximators,
also possess this property, they usually require a larger number of components to achieve
similar approximation accuracy (Barron, 1993). ANNs are thus a parsimonious approach
to nonparametric functional analysis.
ANNs have been widely applied to solve many difficult problems in different areas, including pattern recognition, signal processing, language learning, etc. Since White (1988),
there have also been numerous applications of ANN in economics and finance. Unfortunately, the ANN literature is not easy to penetrate, so it is hard for applied economists to
understand why ANN works and how it can be implemented properly. Fortunately, while
the ANN jargon originated from cognitive science and computer science, they often have
econometric interpretations. For example, a “target” is, in fact, a dependent variable
of interest, an “input” is an explanatory variable, and network “learning” amounts to
1
the estimation of unknown parameters in a network. The purpose of this entry is thus
two-fold. First, this entry introduces ANN using familiar econometric terminology and
hence serves to bridge the gap between the fields of ANN and economics. Second, this
entry provides an overview of ANN modeling approach and its implementation methods.
For an early review of ANN from an econometric perspective, we refer to Kuan and
White (1994).
This entry proceeds as follows. We introduce various ANN model specifications and
the choices of network functions in Section 2. We present the “universal approximation”
property of ANN in Section 3. Model estimation and model complexity regularization
are discussed in Section 4. Section 5 concludes.
2
ANN Model Specifications
Let Y denote the collection of n variables of interest with the t-th observation y t (n × 1)
and X the collection of m explanatory variables with the t-th observation xt (m × 1). In
the ANN literature, the variables in Y are known as targets or target variables, and the
variables in X are inputs or input variables. There are various ways to build an ANN
model that can be used to characterize the behavior of y t using the information contained
in the input variables xt . In this section, we introduce some network architectures and
the functions that are commonly used to build an ANN.
2.1
Feedforward Neural Networks
We first consider a network with an input layer, an output layer, and a hidden layer in
between. The input (output) layer contains m input units (n output units) such that each
unit corresponds to a particular input (output) variable. In the hidden layer, there are q
hidden units connected to all input and output units; the strengths of such connections
are labeled by (unknown) parameters known as the network connection weights. In
particular, γ h = (γh,1 , . . . , γh,m )0 denotes the vector of the connection weights between
the h-th hidden unit and all m input units, and β j = (βj,1 , . . . , βj,q )0 denotes the vector
of the connection weights between the j-th output unit and all q hidden units. An ANN
in which the sample information (signals) are passed forward from the input layer to
the output layer without feedback is known as a feedforward neural network. Figure 1
illustrates the architecture of a 3-layer feedforward network with 3 input units, 4 hidden
units and 2 output units.
2
O1
O2
output layer
activation function F, β weights
hidden layer
activation function G, γ weights
input layer
X1
X2
X3
Figure 1: A feedforward network with 3 input units, 4 hidden units and 2 output units.
This multi-layered structure of a feedforward network is designed to function as a
biological neural system. The input units are the neurons that receive the information
(stimuli) from the outside environment and pass them to the neurons in a middle layer
(i.e., hidden units). These neurons then transform the input signals to generate neural
signals and forward them to the neurons in the output layer. The output neurons in turn
generate signals that determine the action to be taken. Note that all information from
the units in one layer are processed simultaneously, rather than sequentially, by the units
in an “upper” layer.1
Formally, the input units receive the information xt and send to all hidden units,
weighted by the connection weights between the input and hidden units. This information
is then transformed by the activation function G in each hidden unit. That is, the h-th
hidden unit receives x0t γ h and transforms it to G(x0t γ h ). The information generated by
all hidden units is further passed to the output units, again weighted by the connection
weights, and transformed by the activation function F in each output unit. Hence, the
P
j-th output unit receives qh=1 βj,h G(x0t γ h ) and transforms it into the network output:
ot,j = F
q
X
βj,h G(x0t γ h ) ,
j = 1, . . . , n.
(1)
h=1
The output Oj is used to describe or predict the behavior of the j-th target Yj .
In practice, it is typical to include a constant term, also known as the bias term, in
1
This concept, also known as parallel processing or massive parallelism, differs from the traditional
concept of sequential processing and led to a major advance in designing computer architecture.
3
each activation function in (1). That is,
q
X
ot,j = F βj,0 +
βj,h G(γh,0 + x0t γ h ) ,
j = 1, . . . , n,
(2)
h=1
where γh,0 is the bias term in the h-th hidden unit and βj,0 is the bias term in the
j-th output unit. A constant term in each activation function adds flexibility to hiddenunit and output-unit responses (activations), in a way similar to the constant term in
(non)linear regression models. Note that when there is no transformation in the output
units, F is an identity function (i.e., F (a) = a) so that
ot,j = βj,0 +
q
X
βj,h G(γh,0 + x0t γ h ),
j = 1, . . . , n,
(3)
h=1
It is also straightforward to construct networks with two or more hidden layers. For
simplicity, we will focus on the 3-layer networks with only one hidden layer.
While parametric econometric models are typically formulated using a given function
of the input xt , the network (2) is a class of flexible nonlinear functions of xt . The exact
form of a network model depends on the activation functions (F and G) and the number
of hidden units (q). In particular, the network function in (3) is an affine transformation
of G and hence may be interpreted as an expansion with the “basis” function G.
The networks (2) and (3) can be further extended. For example, one may construct
a network in which the input units are connected not only to the hidden units but
also directly to the output units. This leads to networks with shortcut connections.
Corresponding to (2), the outputs of a feedforward network with shortcuts are
q
X
βj,h G(γ h,0 + x0t γ h ) ,
ot,j = F βj,0 + x0t αj +
j = 1, . . . , n,
h=1
where αj is the vector of connection weights between the output and input units. and
corresponding to (3), the outputs are
ot,j = βj,0 + x0t αj +
q
X
βj,h G(γ h,0 + x0t γ h ),
j = 1, . . . , n.
h=1
Figure 2 illustrates the architecture of a feedforward network with 2 input units, 3 hidden
units, 1 output unit and shortcut connections. Thus, parametric econometric models may
be interpreted as feedforward networks with shortcut connections but no hidden-layer
P
connections. The linear combination of hidden-unit activations, qh=1 βj,h G(γ h,0 +x0t γ h ),
in effect characterizes the nonlinearity not captured by the linear function of xt .
4
O1
output layer
activation function F, β weights
hidden layer
activation function G, γ weights
input layer
X1
X2
Figure 2: A feedforward neural network with shortcuts.
2.2
Recurrent Neural Networks
From the preceding section we can see that there is no “memory” device in feedforward
networks that can store the signals generated earlier. Hence, feedforward networks treat
all sample information as “new;” the signals in the past do not help to identify data
features, even when sample information exhibits temporal dependence. As such, a feedforward network must be expanded to a large extent so as to represent complex dynamic
patterns. This causes practical difficulty because a large network may not be easily
implemented. To utilize the information from the past, it is natural to include lagged
target information y t−k , k = 1, . . . , s, as input variables, similar to linear AR and ARX
models in econometric studies. Yet, such networks do not have any built-in structure
that can “memorize” previous neural responses (transformed sample information). The
so-called recurrent neural networks overcome this difficulty by allowing internal feedbacks
and hence are especially appropriate for dynamic problems.
Jordan (1986) first introduced a recurrent network with feedbacks from output units.
That is, the output units are connected to input units but with time delay, so that the
network outputs at time t − 1 are also the input information at time t. Specifically, the
outputs of a Jordan network are
ot,j = F βj,0 +
q
X
βj,h G(γ h,0 + x0t γ h + o0t−1 δ h ) ,
j = 1, . . . , n.
(4)
h=1
where δ h is the vector of the connection weights between the h-th hidden unit and the
input units that receive lagged outputs ot−1 = (ot−1,1 , . . . , ot−1,n )0 . The network (4) can
be further extended to allow for more lagged outputs ot−2 , ot−3 , . . . .
5
O1
O1
output layer
hidden layer
input layer
X1
X1
X2
X2
with delay
(b) Elman network
Figure 3: Recurrent neural networks: Jordan (left) and Elman (right).
(a) Jordan network
Similarly, Elman (1990) considered a recurrent network in which the hidden units are
connected to input units with time delay. The outputs of an Elman network are:
ot,j = F βj,0 +
q
X
βj,h at,h ,
j = 1, . . . , n,
(5)
h=1
at,h = G(γ h,0 + x0t γ h + a0t−1 δ h ),
h = 1, . . . , q,
where at−1 = (at−1,1 , . . . , at−1,q )0 is the vector of lagged hidden-unit activations, and
δ h here is the vector of the connection weights between the h-th hidden unit and the
input units that receive lagged hidden-unit activations at−1 . The network (5) can also
be extended to allow for more lagged hidden-unit activations at−2 , at−3 , . . . . Figure 3
illustrates the architectures of a Jordan network and an Elman network.
From (4) and (5) we can see that, by recursive substitution, the outputs of these recurrent networks can be expressed in terms of current and all past inputs. Such expressions
are analogous to the distributed lag model or the AR representation of an ARMA model
(when the inputs are lagged targets). Thus, recurrent networks incorporate the information in the past input variables without including all of them in the model. By contrast,
a feedforward network requires a large number of inputs to carry such information. Note
that the Jordan network and the Elman network summarize past input information in different ways and hence have their own merits. When the previous “location” of a network
is crucial in determining the next move, as in the design of a robot, a Jordan network
seems more appropriate. When the past internal neural responses are more important,
as in language learning problems, an Elman network may be preferred.
6
2.3
Choices of Activation Function
As far as model specifications are concerned, the building blocks of an ANN model are
the activation functions F and G. Different choices of the activation functions result
in different network models. We now introduce some activation functions commonly
employed in empirical studies.
Recall that the hidden units play the role of neurons in a biological system. Thus, the
activation function in each hidden unit determines whether a neuron should be turned on
or off. Such an on/off response can be easily represented using an indicator (threshold)
function, also known as a heaviside function in the ANN literature, i.e.,
(
1, if γh,0 + x0t γ h ≥ c,
0
G(γh,0 + xt γ h ) =
0, if γh,0 + x0t γ h < c,
where c is a pre-determined threshold value. That is, depending on the strength of
connection weights and input signals, the activation function G will determine whether
a particular neuron is on (G(γh,0 + x0t γ h ) = 1) or off (G(γh,0 + x0t γ h ) = 0).
In a complex neural system, neurons need not have only an on/off response but may
be in an intermediate position. This amounts to allowing the activation function to
assume any value between zero and one. In the ANN literature, it is common to choose a
sigmoid (S-shaped) and squashing (bounded) function. In particular, if the input signals
are “squashed” between zero and one, the activation function is understood as a smooth
counterpart of the indicator function. A leading example is the logistic function:
G(γh,0 + x0t γ h ) =
1
.
1 + exp −[γh,0 + x0t γ h ]
which approaches one (zero) when its argument goes to infinity (negative infinity). Hence,
the logistic activation function generates a partially on/off signal based on the received
input signals.
Alternatively, the hyperbolic tangent (tanh) function, which is also a sigmoid and
squashing function, can serve as an activation function:
exp γh,0 + x0t γ h − exp −[γh,0 + x0t γ h ]
0
.
G(γh,0 + xt γ h ) =
exp γh,0 + x0t γ h + exp −[γh,0 + x0t γ h ]
Compared to the logistic function, this function may assume negative values and is
bounded between −1 and 1. It approaches 1 (−1) when its argument goes to infinity (minus infinity). This function is more flexible because the negative values, in effect,
7
1
1
0.5
0
-1
0
(b) logistic
(c) tanh
Figure 4: Activation functions: logistic (left) and
tanh (right).
represent “suppressing” signals from the hidden unit. See Figure 4 for an illustration of
the logistic and tanh functions. Note that for the logistic function G, a re-scaled function
e such that G(a)
e
G
= 2G(a) − 1 also generates values between −1 and 1 and may be used
in place of the tanh function.2
The aforementioned activation functions are chosen for convenience because they are
differentiable everywhere and their derivatives are easy to compute. In particular, when
G is the logistic function,
dG(a)
= G(a)[1 − G(a)];
da
when G is the tanh function,
2
dG(a)
2
=
= sech2 (a).
da
exp(a) + exp(−a)
These properties facilitate parameter estimation, as will be seen in Section 4.1. Nevertheless, these functions are not necessary for building proper ANNs. For example,
smooth cumulative distribution functions, which are sigmoidal and squashing, are also
legitimate candidates for activation function. In Section 3, it is shown that, as far as
network approximation property is concerned, the activation function in hidden units
does not even have to be sigmoidal, yet boundedness is usually required. Thus, sine and
cosine functions can also serve as an activation function.
As for the activation function F in the output units, it is common to set it as the
identity function so that the outputs of (3) enjoy the freedom of assuming any real
2
A choice of the activation function in classification problems is the so-called radial basis function. We
do not discuss this choice because its argument is not an affine transformation of inputs and hence does
not fit in our framework here. Moreover, the networks with this activation function provide only local
approximation to unknown functions, in contrast with the approximation property discussed in Section 3.
8
value. This choice suffices for the network approximation property discussed in Section 3.
When the target is a binary variable taking the values zero and one, as in a classification
problem, F may be chosen as the logistic function so that the outputs of (2) must fall
between zero and one, analogous to a logit model in econometrics.
3
ANN as an Universal Approximator
What makes ANN a useful econometric tool is its universal approximation property which
basically means that a multi-layered ANN with a large number of hidden units can well
approximate a large class of functions. This approximation property is analogous to that
of nonparametric approximators, such as polynomials and Fourier series, yet it is not
shared by parametric econometric models.
To present the approximation property, we consider the network function element by
element. Let fG,q : Rm × Θm,q → R denote the network function with q hidden units, the
output activation function F being the identity function, and the hidden-unit activation
function G, i.e.,
fG,q (x; θ) = β0 +
q
X
βh G(γh,0 + x0 γ h ),
h=1
as in (3), where Θm,q is the parameter space whose dimension depends on m and q, and
θ ∈ Θm,q (note that the subscripts m and q for θ are suppressed). Given the activation
function G, the collection of all fG,q functions with different q is:
FG =
∞
[
(
fG,q : fG,q (x; θ) = β0 +
q=1
q
X
)
βh G(γh,0 + x0 γ h ) ;
h=1
when the union is taken up to a finite number N , the resulting collection is denoted as
N . Intuitively, F is capable of functional approximation because f
FG
G
G,q can be viewed
as an expansion with the “basis” function G and hence is similar to a nonparametric
approximator.
More formally, we follow Hornik (1991) and consider two measures of the closeness
between functions. First define the uniform distance between functions f and g on the
set K as
dK (f, g) = sup |f (x) − g(x)|.
x∈K
9
Let K denote a compact subset in Rm and C(K) denote the space of all continuous
functions on K. Then, when the activation function G is continuous, bounded and nonconstant, the collection FG is dense in C(K) for all K in Rm in terms of dK (Theorem 2
of Hornik, 1991).3 That is, for any function g in C(K) and any ε > 0, there is a network
N is not dense in C(K) for any
function fG,q in FG such that dK (fG,q − g) < ε. As FG
finite number N , this result shows that any continuous function can be approximated
arbitrarily well on compacta by a 3-layered feedforward network fG,q , provided that q,
the number of hidden units, is sufficiently large.
Taking x as random variables, defined in the probability space with the probability
measure IP, we consider the Lr -norm of f (x) − g(x):
Z
kf − gkr =
1/r
|f (x) − g(x)| d IP(x)
,
r
Rm
1 ≤ r < ∞. For r = 2 (r = 1), this is the well known measure of mean squared error
(mean absolute error). Then, when the activation function G is bounded and nonconstant,
the collection FG is dense in the Lr space (Theorem 1 of Hornik, 1991). That is, any
function g (with finite Lr -norm) can also be well approximated by a 3-layered feedforward
network fG,q in terms of Lr -norm when q is sufficiently large.
It should be emphasized that the universal approximation property of a feedforward
network hinges on the 3-layered architecture and the number of hidden units, but not
on the activation function per se. As stated above, the activation function in the hidden
unit can be a general bounded function and does not have to be sigmoidal. Hornik (1993)
provides results that permit even more general activation functions. Moreover, a feedforward network with only one hidden layer suffices for such approximation property. More
hidden layers may be helpful in certain applications but are not necessary for functional
approximation.
Barron (1993) further derived the rate of approximation in terms of mean squared
error kf − gk22 . It was shown that 3-layered feedforward networks fG,q with G a sigmoidal
function can achieve the approximation rate of order O(1/q), for which the number of
parameters grows linearly with q (with the order O(mq)). This is in sharp contrast
with other expansions, such as polynomial (with p the degree of the polynomial) and
3
Hornik (1991) considered the network without the bias term in the output unit, i.e., β0 = 0. Yet as
long as G is not a constant function, all the results in Hornik (1991) carry over; see Stinchcombe and
White (1998) for details.
10
spline (with p the number of knots per coordinate), which yield suitable approximation
when the number of parameters grows exponentially (with the order O(pm )). Thus, it is
practically difficult for such expansions to approximate well when the dimension of the
input space, m, is large.
4
Implementation of ANNs
In practice, when the activation functions in an ANN are chosen, it remains to estimate its
connection weights (unknown parameters) and to determine a proper number of hidden
units. Given that the connection weights of an ANN model are unknown, this network
must be properly “trained” so as to “learn” the unknown weights. This is why parameter
estimation is referred to as network learning and the sample used for parameter estimation
is referred to a training sample in the ANN literature. As the number of hidden units
q determines network complexity, finding a suitable q is known as network complexity
regularization.
4.1
Model Estimation
The network parameters can be estimated by either on-line or off-line methods. An online learning algorithm is just a recursive estimation method which updates parameter
estimates when new sample information becomes available. By contrast, off-line learning
methods are based on fixed training samples; standard econometric estimation methods
are typically off-line.
To ease the discussion of model estimation, we focus on the simple case that there
is only one target variable y and the network function fG,q . Generalization to the case
with multiple target variables and vector-valued network functions is straightforward.
Once the activation function G is chosen and the number of hidden units is given, fG,q
is a nonlinear parametric model for the target y; the network with multiple outputs is a
system of nonlinear models. Taking mean squared error as the criterion, the parameter
vector of interest θ ∗ thus minimizes
IE[y − fG,q (x; θ)]2 .
(6)
It is well known that
2
2
2
IE y − fG,q (x; θ) = IE y − E(y|x) + IE IE(y|x) − fG,q (x; θ) .
11
As IE(y|x) is the best L2 predictor of y, θ ∗ must also minimize the mean squared approxi
2
mation error: IE IE(y|x)−fG,q (x; θ) . This shows that, among all 3-layered feedforward
networks with the activation function G and q hidden units, fG,q (x; θ ∗ ) provides the best
approximation to the conditional mean function.
Given a training sample of T observations, an estimator of θ ∗ can be obtained by
minimizing the sample counterpart of (6):
T
1X
[yt − fG,q (xt ; θ)]2 ,
T
t=1
which is just the objective function of the nonlinear least squares (NLS) method. The
NLS method is an off-line estimation method because the size of the training sample
is fixed. Under very general conditions on the data and nonlinear function, it is well
known that the NLS estimator is strongly consistent for θ ∗ and asymptotically normally
distributed; see, e.g., Gallant and White (1988).
In many ANN applications (e.g., signal processing and language learning), the training sample is not fixed but constantly expands with new data. In such cases, off-line
estimation may not be feasible, but on-line estimation methods, which update the parameter estimates based solely on the newly available data, are computationally more
tractable. Moreover, on-line estimation methods can be interpreted as “adaptive learning” by biological neural systems. It should be emphasized that when there is only a
given sample, as in most empirical studies in economics, recursive estimation is not to
be preferred because it is, in general, statistically less efficient than the NLS method in
finite samples.
Note that the parameter of interest θ ∗ is the zero of the first order condition of (6):
IE ∇fG,q (x; θ) y − fG,q (x; θ) = 0,
where ∇fG,q (x; θ) is the (column) gradient vector of fG,q with respect to θ. To estimate
θ ∗ , a recursive algorithm proposed by Rumelhart, Hinton, and Williams (1986) is
θ̂ t+1 = θ̂ t + ηt ∇fG,q (xt ; θ̂ t ) yt − fG,q (xt ; θ̂ t ) ,
(7)
where ηt > 0 is a parameter that re-scales the adjustment term in the square bracket.
It can be seen from (7) that the adjustment term is determined by the gradient descent
12
direction and the error between the target and network output: yt − fG,q (xt ; θ̂ t ), and it
requires only the information at time t, i.e., yt , xt , and the estimate θ̂ t .4
The algorithm (7) is known as the error back-propagation (or simply back-propagation)
algorithm in the ANN literature, because the error signal [yt − fG,q (xt ; θ̂ t )] is propagated
back through the network to determine the change of each weight. The underlying idea
of this algorithm can be traced back to the classical stochastic approximation method introduced in Robins and Monro (1951). White (1989) established consistency and asymptotic normality of θ̂ t in (7). Note that the parameter ηt in the algorithm is known as
P
a learning rate. For consistency of θ̂ t , it is required that ηt satisfies ∞
t=1 ηt = ∞ and
P∞ 2
t=1 ηt < ∞, e.g., ηt = 1/t. The former condition ensures that the updating process
may last indefinitely, whereas the latter implies ηt → 0 so that the adjustment in the
parameter estimates can be made arbitrarily small.5
Instead of the gradient descent direction, it is natural to construct a recursive algorithm with a Newton search direction. Kuan and White (1994) proposed the following
algorithm:
0
c
c
c
H
t+1 = H t + ηt ∇fG,q (xt ; θ̂ t )∇fG,q (xt ; θ̂ t ) − H t ,
c −1 ∇f (x ; θ̂ ) y − f (x ; θ̂ ) ,
θ̂ t+1 = θ̂ t + ηt H
t+1
G,q
t t
t
G,q
t t
(8)
c
where H
t+1 characterizes a Newton direction and is recursively updated via the first
√
equation. Kuan and White (1994) showed that θ̂ t in (8) is t-consistent, statistically
more efficient than θ̂ t in (7), and asymptotically equivalent to the NLS estimator. The
algorithm (8) may be implemented in different ways; for example, there is an algorithm
that is algebraically equivalent to (8) but does not involve matrix inversion. See Kuan
and White (1994) for more discussions on the implementation of the Newton algorithms.
On the other hand, estimating recurrent networks is more cumbersome. From (4) and
(5) we can see that recurrent network functions depend on θ directly and also indirectly
through the presence of internal feedbacks (i.e., lagged output and lagged hidden-unit
activations). The indirect dependence on parameters must be taken into account in
4
The algorithm (7) is analogous to the numerical steepest-descent algorithm. However, (7) utilizes
only the information at time t, whereas numerical optimization algorithms are computed using all the
information in a given sample and hence are off-line methods.
5
In many applications of ANN, the learning rate is often set to a constant ηo ; the resulting estimate θ̂t
loses consistency in this case. Kuan and Hornik (1991) established a convergence result based on small-ηo
asymptotics.
13
calculating the derivatives with respect to θ. Thus, NLS optimization algorithms that
require analytic derivatives are difficult to implement. Kuan, Hornik, and White (1994)
proposed the dynamic back-propagation algorithm for recurrent networks, which is analogous to (7) but involves more updating equations. Kuan (1995) further proposed a
√
Newton algorithm for recurrent networks, analogous to (8), and showed that it is tconsistent and statistically more efficient than the dynamic back-propagation algorithm.
We omit the details of these algorithms; see Kuan and Liu (1995) for an application of
these estimation methods for both feedforward and recurrent networks.
Note that the NLS method and recursive algorithms all require computing the derivatives of the network function. Thus, a smooth and differentiable activation function, as
the examples given in Section 2.3, are quite convenient for network parameter estimation.
Finally, given that ANN models are highly nonlinear, it is likely that there exist multiple
optima in the objective function. There is, however, no guarantee that the NLS method
and the recursive estimation methods discussed above will deliver the global optimum.
This is a serious problem because the dimension of the parameter space is typically large.
Unfortunately, a convenient and effective method for finding the global optimum in ANN
estimation is not yet available.
4.2
Model Complexity Regularization
Section 3 shows that a network model fG,q can approximate unknown function when the
number of hidden units, q, is sufficiently large. When there is a fixed training sample,
a complex network with a very large q may over fit the data. Thus, there is a trade-off
between approximation capability and over-fitting in implementing ANN models.
An easy approach to regularizing the network complexity is to apply a model selection criteria,6 such as Schwarz (Bayesian) information criterion (BIC), to the network
models with various q. As is well known, BIC consists of two terms: one is based on
model fitness, and the other penalizes model complexity. Hence, it is suitable for regularizing network complexity; see also Barron (1991). A different criterion introduced in
Rissanen (1986, 1987) is predictive stochastic complexity (PSC) which is just an average
6
Alternatively, one may consider testing whether some hidden units may be dropped from the model.
This amounts to testing, say, βh = 0 for some h. Unfortunately, the parameters in that hidden-unit
activation function (γh,0 and γ h ) are not identified under this null hypothesis. It is well known that,
when there are unidentified nuisance parameters, standard econometric tests are not applicable.
14
of squared prediction errors:
PSC =
T
X
2
1
yt − fG,q (xt , θ̂ t ) ,
T −k
t=k+1
where θ̂ t is the predicted parameter estimate based on the sample information up to
time t − 1, and k is the total number of parameters in the network. Given the number
of inputs, the network with the smallest BIC or PSC gives the desired number of hidden
units q ∗ . Rissanen showed that both BIC and PSC can be interpreted as the criteria for
“minimum description length,” in the sense that they determine the shortest code length
(asymptotically) that is needed to encode a sequence of numbers. In other words, these
criteria lead to the least complex model that still captures the key information in data.
Swanson and White (1997) showed that a network selected by BIC need not perform well
in out-of-sample forecasting, however.
Clearly, PSC requires estimating the parameters at each t. It would be computationally demanding if the NLS method is to be used, even for a moderate sample. For
simplicity, Kuan and Liu (1995) suggested a two-step procedure for implementing ANN
models. In the first step, one estimates the network models and computes the resulting
PSCs using the recursive Newton algorithm, which is asymptotically equivalent to the
NLS method. When a suitable network structure is determined, the Newton parameter estimates can be used as initial values for NLS estimation in the second step. This
approach thus maintains a balance between computational cost and estimator efficiency.
5
Concluding Remarks
In this entry, we introduce ANN model specifications, their approximation properties,
and the methods for model implementation from an econometric perspective. It should
be emphasized that ANN is neither a magical econometric tool nor a “black box” that can
solve any difficult problems in econometrics. As discussed above, a major advantage of
ANN is its universal approximation property, a property shared by other nonparametric
approximators. Yet compared with parametric econometric models, a simple ANN need
not perform better, and a more complex ANN (with a large number of hidden units) is
more difficult to implement properly and can not be applied when there is only a small
data set. Therefore, empirical applications of ANN models must be exercised with care.
15
References
Barron, A. R. (1991). Complexity regularization with application to artificial neural
networks, in Nonparametric Functional Estimation and Related Topics, G. Roussas
(ed.), pp. 561–576, Dordrecht, the Netherlads: Kluwer Academic Publishers.
Barron, A. R. (1993). Universal approximation bounds for superpositions of a sigmoidal
function, IEEE Transactions on Information Theory, IT-39, 930–945.
Cybenko, G. (1989). Approximation by superpositions of a sigmoid function, Mathematics of Control, Signals, and Systems, 2, 303–314.
Elman, J. L. (1990). Finding structure in time, Cognitive Science, 14, 179-211.
Funahashi, K. (1989). On the approximate realization of continuous mappings by neural
networks, Neural Networks, 2, 183–192.
Gallant, A. R. and H. White (1988). A Unified Theory of Estimation and Inference for
Nonlinear Dynamic Models, Oxford, UK: Basil Blackwell.
Hornik, K. (1991). Approximation capabilities of multilayer feedforward nets, Neural
Networks, 4, 231–242.
Hornik, K. (1993). Some new results on neural network approximation, Neural Networks,
6, 1069–1072.
Hornik, K., M. Stinchcombe, and H. White (1989). Multi-layer feedforward networks are
universal approximators, Neural Networks, 2, 359–366.
Hornik, K., M. Stinchcombe, and H. White (1990). Universal approximation of an unknown mapping and its derivatives using multilayer feedforward networks, Neural
Networks, 3, 551–560.
Jordan, M. I. (1986). Serial order: A parallel distributed processing approach, Institute
for Cognitive Science Report 8604, UC San Diego.
Kuan, C.-M. (1995) “A recurrent Newton algorithm and its convergence properties”,
IEEE Transactions on Neural Networks, 6, 779–783.
Kuan, C.-M. and K. Hornik (1991). Convergence of learning algorithms with constant
learning rates, IEEE Transactions on Neural Networks, 2, 484–489.
Kuan, C,-M., K. Hornik, and H. White (1994). A convergence result for learning in
recurrent neural networks, Neural Computation, 6, 420–440.
16
Kuan, C.-M. and T. Liu (1995). Forecasting exchange rates using feedforward and recurrent neural networks, Journal of Applied Econometrics, 10, 347–364.
Kuan, C.-M. and H. White (1994). Artificial neural networks: An econometric perspective (with discussions), Econometric Reviews, 13, 1–91 and 139–143.
McClelland, J. L., D. E. Rumelhart, and the PDP Research Group (1986). Parallel
Distributed Processing: Explorations in the Microstructure of Cognition, Vol. 2,
Cambridge, MA: MIT Press.
McCulloch, W. S. and W. Pitts (1943). A logical calculus of the ideas immanent in
nervous activity. Bulletin of Mathematical Biophysics, 5, 115–133.
Rissanen, J. (1986). Stochastic complexity and modeling, Annals of Statistics, 14, 1080–
1100.
Rissanen, J. (1987). Stochastic complexity (with discussions), Journal of the Royal Statistical Society, B, 49, 223–239 and 252–265.
Robbins, H. and S. Monro (1951). A stochastic approximation method, Annals of Mathematical Statistics, 22, 400–407.
Rosenblatt, F. (1958). The perception: A probabilistic model for information storage
and organization in the brain, Psychological Reviews, 62, 386–408.
Rumelhart, D. E., G. E. Hinton, and R. J. Williams (1986). Learning internal representation by error propagation, in Parallel Distributed Processing: Explorations in the
Microstructure of Cognition, Vol. 1, pp. 318–362, Cambridge, MA: MIT Press.
Rumelhart, D. E., J. L. McClelland, and the PDP Research Group (1986). Parallel
Distributed Processing: Explorations in the Microstructure of Cognition, Vol. 1,
Cambridge, MA: MIT Press.
Stinchcombe, M. B. and H. White (1998). Consistent specification testing with nuisance
parameters present only under the alternative, Econometric Theory, 14, 295–325.
Swanson, N. R. and H. White (1997). A model selection approach to real-time macroeconomic forecasting using linear models and artificial neural networks, Review of
Economics and Statistics, 79, 540–550.
White, H. (1988). Economic prediction using neural networks: The case of IBM stock
prices. Proceedings of the Second Annual IEEE Conference on Neural Networks,
pp. II:451–458. New York: IEEE Press.
17
White, H. (1989). Some asymptotic results for learning in single hidden layer feedforward
network models, Journal of the American Statistical Association, 84, 1003–1013.
18
Download